Reaction of NO with the Reduced R2 Protein of Ribonucleotide

Mar 31, 1995 - The active R2 protein of ribonucleotide reductase from Escherichia coli contains a catalytically essential tyrosine radical at position...
0 downloads 0 Views
Biochemistry 1995, 34, 11090- 11098

11090

Reaction of NO with the Reduced R2 Protein of Ribonucleotide Reductase from Escherichia coli? Charlene J. Haskin,* Natarajan Ravi,§ John B. Lynch,$." Eckard Munck,**§and Lawrence Que, Jr.*,t Department of Chemistry, University of Minnesota, Minneapolis, Minnesota 55455, and Department of Chemistry, Carnegie Mellon University, Pittsburgh, Pennsylvania 15213 Received March 31, 1995; Revised Manuscript Received July 10, 1995@

ABSTRACT: The active R2 protein of ribonucleotide reductase from Escherichia coli contains a catalytically essential tyrosine radical at position 122 (Ty1-122.) that is formed during the reaction of dioxygen with the nearby diiron(I1) center. To gain insight into the mode of dioxygen binding, the reaction of the 0 2 analog NO with the diiron(I1) centers of R2,d has been investigated by spectroscopic methods. R2,ed reacts with NO to form an adduct with visible absorption features at 450 and 620 nm and Mossbauer parameters (6 = 0.75 "/s, AEQ = -2.13 and -1.73 "/s) typical of those observed for S = 3/2 {FeN0}7 complexes of other non-heme iron proteins. However, unlike other non-heme {FeN0}7 complexes, this adduct is EPR silent. Our Mossbauer studies show that each iron site of R2,ed binds one NO to form local S = 3/2 {FeN0}7 centers which then couple antiferromagnetically (J RZ 5 cm-', 5T=JSI-S~) to afford an {FeN0}2 center (77% of total iron). This {FeN0}2 center decomposes with a first-order rate constant of 0.013 min-' to form R2,,,, accompanied by the release of N20. These observations suggest that both iron(I1) ions of the two diiron(I1) centers of R2red have available sites for NO binding, in agreement with the crystallographic results on R2red, and that the bound NO molecules are sufficiently close to each other to permit N-N bond formation to produce N20. These observations support the proposal that dioxygen binding may also involve both metal ions of the diiron(I1) center to form a (p-l,l-,or p-1,2-peroxo)diiron(II1) center. This observed reactivity of R2,d with NO may contribute to the in viuo inhibition of ribonucleotide reductase by NO.

The enzyme ribonucleotide reductase catalyzes the conversion of ribonucleotides to deoxyribonucleotides and provides a balance of each of the four deoxyribonucleotides needed for the first step of DNA biosynthesis (Reichard, 1993; Stubbe, 1990). Ribonucleotide reductase from Escherichia coli is composed of two proteins, R1 and R2. The R1 protein contains binding sites for substrates and allosteric effectors and provides the reducing equivalents for substrate reduction (Uhlin & Eklund, 1994). The R2 protein in the enzymatically active form, R20x,1contains a (,u-oxo)diiron(III)center and a stable tyrosine radical at position 122 (Tyr122.) (Larsson & Sjoberg, 1986) positioned about 5 away from the nearest iron (Nordlund & Eklund, 1993; Nordlund et al., 1990). This radical initiates ribonucleotide reduction on the R1 protein, presumably by long-range transfer of this oxidizing equivalent (Nordlund & Eklund, 1993). The formation of the essential Tyr122' occurs during the reaction of dioxygen with the diiron(I1) form of the protein,

A

This work was supported by grants from the National Science Foundation (MCB-9405723, L.Q.) and the National Institutes of Health (GM22701, E.M.) and by a National Institutes of Health predoctoral traineeship for C.J.H. (GM08277). * To whom correspondence should be addressed. University of Minnesota. QCarnegieMellon University. 'I Current address: W. R. Grace & Co.-Conn., Washington Research Center, 7379 Route 32, Columbia, MD 21044. Abstract published in Advance ACS Abstracts, August 15, 1995. Abbreviations: R2,,, oxidized R2 protein of ribonucleotide reductase: R2,,,, tyrosine radical-reduced diiron(II1) form of the R2 protein of ribonucleotide reductase: R2,d, reduced R2 protein of ribonucleotide reductase: GC, gas chromatography: 3,4-PCD, protocatechuate 3,4dioxygenase; 4,5-PCD, protocatechuate 4,5-dioxygenase; TPA, tris(2pyridylmethy1)amine.

*

'

@

0006-2960/95/0434-11090$09.00/0

R2,,d. Recent X-ray crystallographic studies of the S211A mutant protein (S211 is not a ligand to the diiron center) suggests that the diiron(I1) centers of R2,d consist of two ions, each ligated to one terminal histidine, one terminal carboxylate, and two bridging carboxylate ligands (Aberg, 1993). Some mechanisms that have been proposed for the TyrIz2*radical generation reaction (Ling et al., 1994b; Que, 1991) invoke a high valent iron intermediate, by analogy to those found in heme-containing enzymes (Dawson, 1988; Guengerich, 1991). A similar mechanism has also been proposed for the non-heme diiron center of methane monooxygenase (Lee et al., 1993). However, the recent characterization of intermediates in the radical generation mechanism of R2 has determined that the oxidizing equivalents do not necessarily reside on the iron centers but instead reside on a metal ligand or nearby protein residue (Bollinger et al., 1994a,b; Ravi et al., 1994). The differences observed in dioxygen reactivity at heme and non-heme iron centers provide an example of the variety of ways protein active sites are engineered to accommodate high valent states to accomplish similar chemistry. The recently characterized intermediates in the radical generation reaction of R2 represent species subsequent to 0-0 bond cleavage; thus a peroxide intermediate has not yet been observed, which is presumably the first step of the mechanism (Ling et al., 1994b; Que, 1991). Support for this intermediate has been provided by the observation that H202 reacts with the radical-free (,u-oxo)diiron(III) form of the protein, R2,,, to generate Ty1-122' in a peroxide shunt pathway (Fontecave et al., 1990; Sahlin et al., 1990). Further support has recently arisen from the quantitative incorporation of l80

0 1995 American Chemical Society

Reaction of NO with Ribonucleotide Reductase into the p-oxo bridge of R2,, in the reaction of R2,,d with (Ling et al., 1994b). However, the peroxide coordination mode in this intermediate is not known. A number of proteins with dinuclear metal sites bind dioxygen to form a peroxide intermediate. Hemerythrin, a non-heme diiron protein involved in dioxygen transport, binds dioxygen to form a terminal, $ peroxide (Shiemke et al., 1986; Stenkamp, 1994). Hemocyanin, a dicopper dioxygen transport protein, binds dioxygen to form a p-$:q2 peroxide (Ling et al., 1994a; Magnus et al., 1994). Still other peroxide coordination modes have been observed by a number of diiron and dicopper model complexes (Paul et al., 1991; Dong et al., 1993; Feig & Lippard, 1994; Kitajima & Morooka, 1994; Kitajima et al., 1994). While it is not clear what will govem dioxygen reactivity, it is nevertheless important to gain insight into the dioxygen coordination mode and to relate it to protein function. Valuable information regarding the coordination of dioxygen in the putative diiron peroxide intermediate in R2 may be provided by exploring the reaction of R2,d with nitric oxide. Nitric oxide has been a valuable probe for the study of many non-heme iron proteins (Arciero et al., 1983; Chen et al., 1989; Orville & Lipscomb, 1993). Because of its similar structure and electronic prqperties, it can be used as an 0 2 analog to study the mode of 0 2 binding. For example, deoxyhemerythrin was shown to bind NO reversibly to one iron(I1) site in a bent, end-on coordination mode, analogous to the 02 binding mode in oxyhemerythrin (Nocek et al., 1988; Stenkamp, 1994). In addition, the unpaired electron of NO makes it a convenient spectroscopic probe because NO adducts formed by complexation with high spin Fe(I1) are transformed into S = 3/2 systems that exhibit characteristic EPR signals and Mossbauer spectra (Arciero et al., 1983; Bill et al., 1985; Chen et al., 1989; Brown et al., 1995); these NO complexes have been designated as {FeN0I7 complexes (Enemark & Feltham, 1974). The rich chemistry precedented in the NO complexes of non-heme iron proteins has prompted an examination of the effect of NO on R2,,d.

EXPERIMENTAL PROCEDURES Purification of R2, and j7Fe-Enriched R2,,. R2,, and 57Fe-enriched R2,, were purified from E. coli strain N6405l pSPS2 (Salowe & Stubbe, 1986) according to published procedures (Lynch, 1989; Salowe, 1987). 57Fe-enrichedR2,, was prepared by adding a freshly made 57Fe(II) solution buffered with 1.0 M H E P E N 2 5 mM ascorbic acid (pH 5) to the cell suspension after treatment with lysozyme. A 15min delay after lysozyme addition was found to be essential for achieving a high incorporation (nearly 100%)of 57Feinto the R2 active sites. The purity of R2,, and 57Fe-enriched R2,, was determined by absorption ratios of A28dA410 = 16, A410/A4o5 = 1.19, and A28dA370 = 13.5 and the appearance of a single band on SDS-polyacrylamide gel electrophoresis. The protein concentration was determined by €280 = 141 mM-l cm-' (Lynch, 1989). After purification, R2,, and 57Fe-enriched R2,, were treated identically. Preparation of R2,,, and R2,,d. R2,,, was prepared by adding hydroxyurea to R2,, (0.5 p L of 1.5 M hydroxyurea per mg) and incubating at room temperature for 15 min. Excess hydroxyurea was removed by dialysis for 8 h at 4 "C versus four changes of 50 mM HEPES buffer, pH 7.6, containing 5% glycerol (Barlow et al., 1983). The concen-

Biochemistry, Vol. 34, No. 35, 1995 11091

tration of R2met was determined by €280 = 130 mM-I cm-I. To prepare RZred, a solution of R2,,, and benzyl viologen in a molar ratio of 1:OS was degassed on ice by vacuudargon cycles on a dual-manifold vacuum line equipped with a BASF Catalyst column to remove trace amounts of 0 2 and was then treated with an anaerobic solution of sodium dithionite which immediately turned the protein solution purple due to the presence of reduced benzyl viologen. A dithionite/protein ratio of 4: 1 was used for UV-vis, EPR, and Mossbauer experiments, and a molar ratio of 2:l was used for GC quantitation experiments. Preparation of R2,d NO Samples. R2,,, (0.38-2.0 mL of 0.10-0.50 mM) was reduced and incubated for 3 min in an anaerobic cuvette for UV-vis, EPR, and Mossbauer experiments or in a septum-capped 1-dram vial for GC quantitation experiments. Nitric oxide (in a fume hood) purified by passing through concentrated KOH was anaerobically delivered over the surface and bubbled directly into the protein sample for 1.0-1.5 min. The purple color of reduced benzyl viologen disappeared within approximately 10 s of NO addition, and the yellow-green color of the NO complex developed over the next 30-60 s in the presence of a large excess of NO. EPR and Mossbauer samples were prepared from the same solution by the anaerobic transfer of 300 y L to an EPR tube or Mossbauer cup and frozen in liquid Nz. Physical Measurements. Optical spectra and kinetic data were recorded on a Beckman DU 640 spectrophotometer using 1.0 or 0.2 cm anaerobic cuvettes. EPR spectra were recorded on a Varian E109 spectrometer equipped with an Oxford ESRlO cryostat under the following conditions: temperature, 10 K; microwave frequency, 9.23 GHz; microwave power, 20 p W ; modulation frequency, 100 kHz; and modulation amplitude, 10 G. EPR quantitations were carried out by double integration under nonsaturating conditions and comparison with a 0.474 mM Cu"(EDTA) standard. Mossbauer spectra were recorded on either a strong-field or a weak-field spectrometer operating in a constant acceleration mode in a transmission geometry using Janis Research Inc. cryostats that allow for a variation in temperature from 1.5 to 300 K. One of the dewars housed a superconducting magnet that allowed for the application of magnetic fields up to 8 T parallel to the y-radiation. The zero velocity of the Mossbauer spectra were referred to the centroid of the room-temperature spectrum of a metallic iron foil. Kinetics of { FeNOj2 Complex Decomposition. Samples of R2,,d NO were prepared in an anaerobic cuvette (0.5 mL of 0.40 or 0.50 mM). Kinetic data were collected after the {FeN0}2 complex began to decay by following the 620 nm band. The data were analyzed by two methods, the Kezdy-Swinboume method [Pt = P,[ 1 - exp(kt)l PI+,exp(kt)] and the Guggenheim method [Pt = (Po Pm) [ l - exp(-kt)] exp(-kt)] (Espenson, 1981), both of which are useful in determining the rate constant of a firstorder reaction with no endpoint data. In both methods, the plotted data gave linear fits from which a rate constant was obtained. NzO Quantitation. The separation and detection of N20 was performed on a Hewlett Packard 5890 series I1 gas chromatograph with a HP 3396 series I1 integrator using a Porpak Q column (6 ft) using a 20 mL/min flow rate at 30 "C with helium as the carrier gas. All samples were prepared in argon-flushed septum-capped 1-dram vials and contained

+

+

+

Haskin et al.

11092 Biochemistry, Vol. 34, No. 35, 1995 2.0

I

I

0.28

I

I

I

A

1.6

8 i

1.2

0 12

f!

B 9

0.8-

0.4

-

0.0’

I

I

I

I

400

500

600

700

I 800

Wavelength (nm)

FIGURE1: Optical spectra of 2.0 mL of 0.29 mM R2,,d (50 mM HEPES, pH 7.6, 5% glycerol) plus NO at 5 , 10, 18, 24, and 30 min after initial NO exposure; the arrows indicate the reaction progress. 1.0 mL of 0.50 mM of the NO complex for sample vials or 1.0 mL of buffer for standard and control vials to correct for the volume and gas solubility differences. To each standard vial, 20,30,40,60, or 80 p L of N20 gas was added and a headspace aliquot (30 pL) was injected into the GC from which a standard curve of pmol of N20 vs integration area was prepared. In protein sample vials, a headspace aliquot was injected 3- 10 min after NO addition and again after 87- 105 min. N20 gas was detected at a retention time of 3.7-4.1 min as a well-resolved peak in all samples. The net amount of N20 produced @mol) during this time was determined after comparison with the standard curve. Control vials were treated identically to sample vials; each control vial containing dithionite was treated with NO and analyzed at the same two time points as the protein samples. Careful treatment of control vials was necessary to determine the background net N20 production which was used to correct the net N20 production in the protein samples.

RESULTS Reaction of R2,d with NO. The addition of nitric oxide to an anaerobic solution of R2,d results in NO complexation. Complex formation is highly dependent upon the sample volume and protein concentration; an increase in either decreases the rate of formation, suggesting that the time required for NO dissolution is rate limiting. Because of this complication, the rate constant for the formation of this complex was not determined. NO complex formation can

be observed for a range of volumes (0.38-2.0 mL) and protein concentrations (0.10-0.50 mM); however, the use of protein concentrations above 0.50 mM yields a mixture of R2,,d and the NO complex. To ensure complete reaction with R2,d, excess dissolved NO was always present in the protein solution. The use of a large sample volume (2.0 mL of 0.29 mM R2,d) significantly slowed the reaction between R2,d and NO to reveal the development of the optical spectrum from 5 to 30 min after initial NO exposure (Figure 1). The spectrum taken at 30 min shows maximal formation of this complex and exhibits absorption maxima at 350, 450, and 620 nm. The 450 and 620 nm bands of this yellow-green complex originate from an iron-nitrosyl complex. Similar absorption features, characteristic of iron-nitrosyl complexes, have been observed in a number of mononuclear and dinuclear non-heme iron-nitrosyl complexes in proteins and model complexes (Table 1). Using a variety of spectroscopic techniques, Solomon, Hodgson, and co-workers have recently proposed that the electronic structure of the {FeN0}7 complexes is best described as high-spin iron(II1) antiferromagnetically coupled to NO- (Brown et al., 1995). These researchers have assigned the bands observed around 450 and 620 nm to NO--to-Fe(II1)charge transfer transitions. The 350 nm band observed in the R2,,d NO sample does not originate from the iron-nitrosyl complex, since it is also present in anaerobic samples of R2,, to which NO has been added. This band has previously been assigned in isopenicillin N synthase-NO as arising from a nonspecific interaction of NO with proteins (Chen et al., 1989). Samples of R%d NO, for various protein concentrations, were frozen at different times after initial NO exposure (5100 min) and studied by EPR spectroscopy. The spectra contained three distinct components that had, for different preparations, nearly identical shapes and relative concentrations (Figure 2 ) . The major signal at g = 4.06, 3.98, and 2.01 belongs to an S = 3/2 species indicative of an {FeN0}7 complex (Table 1). Determination of the spin concentration of this signal for three independent R2,d 4-NO preparations showed that it represents only 10-20% of the iron; these quantitations include a minority S = 3/2 species discernible by the weaker signals in the wings of the g 4 feature. A second signal, with a spin concentration of -1-2% of the iron, is observed at g = 2.02 and 2.03. This signal showed slight broadening when 56Fewas replaced by 57Fe,suggesting that it results from an iron-containing complex. Such signals are commonly observed for [Fe(N0)2(SR)2] complexes (Kennedy et al., 1993), and the signal observed here may

+

+

~~

Table 1: Spectroscopic Properties of Non-Heme {FeNO}’ Complexes sample

I,,, (nm) E (cm-’ M-I)

{FeN0}2 complex of R2

450 (760)b 620 (220)b

{FeNO}’ complex of R2 isopenicillin N synthase-NO Fe(EDTA)NO 4,5-PCD-NO putidamonooxin-NO reduced 3,4-PCD-NO soybean lipoxygenase-NO deoxyhemerythrin-NO

450 (760)b 620 (220)b 430 (800) 600 440 (550) 660 (1 10)

(S = l/2)

430 (1870) 650 408 (1200) 600 (500) 500 (700)

6 ( m d s ) AEQ ( d s ) g values Alsoa(MHz) 0.75 -2.13 -31.5 0.75 -1.73 -34.2 0.70 -1.7 4.06, 3.98, 2.01 -34.7 0.75 -1.0 4.09, 3.95, 2.00 -37.0 0.66 -1.67 4.1, 3.9, 2.0 -35.7 0.66 -1.67 4.09, 3.91, 2.0 -32.0 0.68 - 1.4 4.0, 4.0, 2.0 -32 4.34, 3.69, 1.98 4.0, 4.0, 2.0 0.68 0.61 2.77, 1.84, 1.84 1.21 2.65

+ +

reference this work this work Chen et al. (1989) Arciero et al. (1983), Brown et al. (1995) Arciero et al. (1983) Bill et al. (1985) Orville and Lipscomb (1993) Nelson (1987) Nocek et al. (1988)

Isotropic part of 57Fe magnetic hyperfine tensor; Aiso = (A, A, AJ3. Both {FeN0}2 and (FeNO)’ are assumed to contribute with identical E to the optical spectrum which is scaled to reflect 90% of the iron in the sample.

Biochemistry, Vol. 34, No. 35, 1995 11093

Reaction of NO with Ribonucleotide Reductase

I

‘4.06

I-



,97

I

I

I

I

,

I

1000

1500

2000

2500

3000

3500

1

0

H (Gauss)

FIGURE2: X-band EPR spectrum of 0.29 mM R2,,d (50 mM HEPES, pH 7.6,5% glycerol) plus NO frozen 100 min after initial NO exposure. Instrumental conditions: temperature, IO K; microwave frequency, 9.23 GHz; microwave power, 0.02 mW; modulation frequency, 100 kHz; and modulation amplitude, IO G. arise from a similar complex with any of the 10 available solvent-accessible cysteine residues of R2 (Nordlund & Eklund, 1993). A third signal is observed at g x 1.97. This feature was also present in control samples containing buffer NO or R2,,, NO and is attributed to NO in solution. Finally, studies of R2,,d NO in a bimodal cavity did not reveal any integer-spin EPR signal. Thus, our studies show that only a small fraction of the iron in R2,,d iNO samples resides in EPR-active complexes; moreover, none of the signals observed represents a dinuclear center. In order to investigate the EPR-silent species, Mossbauer samples of R2,d NO were obtained under conditions to ensure maximum NO complex formation by following the time course of the optical spectra. The 4.2 K Mossbauer spectrum of a representative sample, frozen 12 min after NO exposure, is shown in Figure 3. The spectrum contains the contribution of three distinct iron complexes (Table 2). The contribution of the S = 3/2 {FeN0}7 complex is readily recognized by the presence of a component that exhibits magnetic hyperfine structure. This component has spectral parameters very similar to those reported for other (FeN0I7 complexes (Table 1); thus, its Mossbauer spectrum could be readily computed from a suitable S = 3/2 spin Hamiltonian (see below); a theoretical curve for this species, scaled to represent 13% of total iron, is drawn through the data of Figure 3. A second species, accounting for 77% of the iron, gives rise to the major doublet in the central portion of the spectrum. We will argue below that it belongs to an EPRsilent dinuclear center which has one NO bound to each iron site. A third, minor component exhibiting a quadrupole doublet with AEQ % 1.4 m d s and 6 % 0.07 m d s appears also in the central portion of the spectrum. This doublet seems to belong to a diamagnetic iron environment of unknown nature; its contribution, representing 10% of the iron, is shown above the data in Figure 3. Five independent preparations of R2,,d NO showed the same collection of species with similar relative concentrations. A control sample, made by anaerobic addition of NO to R2met,exhibited only the two quadrupole doublets (data not shown) characteristic of the native diiron(II1) center of R2,,,, showing that NO does not bind to the diiron(II1) center. The Mossbauer spectrum of the diiron(II1) center of native R2 consists of two quadrupole doublets with 61 = 0.45 m d s

+

+

+

+

+

-10

-8

-6

-4

-2

0

2

4

6

8

10

VELOCITY ( d s )

FIGURE 3: Mossbauer spectrum of 0.48 mM 57Fe-enrichedR2,,d (50 mM HEPES, pH 7.6, 5% glycerol) -I- NO frozen 12 min after exposure to NO. The spectrum was recorded at 4.2 K in a parallel field of 0.05 T. The solid line drawn through the data is a theoretical curve representing the mononuclear S = 3/2 {FeNO}?complex (13% of Fe). The curve was generated from eqs 1-3 using D = 9 cm-l, E/D = 0.02, A = -34.2, -31.5, and -38.4 MHz, AEQ = -1.7 m d s , y = 0, and 6 = 0.70 m d s . The doublet shown above the data indicates the contribution (10% of Fe) of the unidentified species with AEQ = 1.4 mm/s and 6 = 0.04 m d s . Table 2: Mossbauer Parameters and Assignments for Reduced R2 NO Samples

+

fraction (%)

assignment 12 min 100 min 300 min 6 ( m d s ) AEo ( m d s ) 22 0.75,0.75 -2.13, -1.73 {FeN0}2 77 30 {FeNO}’ R2mt unknown

13 0 10

13 42

13 47

14

18

0.70

-1.7

0.55,0.45 - 1.62,-2.44 0.07 1.4

and 62 = 0.55 m d s , while that of R2,d corresponds to two nearly identical high-spin iron(I1) sites with 61 % 62 = 1.25 m d s (Lynch et al., 1989). In contrast, {FeN0}7 complexes have 6 values ranging from 0.66 to 0.75 m d s (Table 1); the {FeNO}’ complex observed here has a similar shift, namely, 6 % 0.70 m d s . A similar shift, 6 = 0.75 m d s , is also observed for the major doublet in the spectrum of Figure 3. A detailed analysis of our data set suggests that this doublet actually consists of two components with 61 x 62 % 0.75 m d s and AEo(1) = -1.73 m d s and A E Q ( ~= ) -2.13 m d s . Since this species produces quadrupole doublets at 4.2 K, rather than magnetically split spectra, the complex must have integer electronic spin. Taken together, our data suggest that we are dealing with a diiron center containing two exchange-coupled S = 3/2 {FeN0}7 sites in an {FeN0}2 complex. This interpretation is supported by the observation that the optical spectra of R2,,d f NO exhibit bands characteristic of {FeNO}? complexes (Table 1). If all of the absorption were attributed to the minority {FeN0}7 complex, this species would have an E x 6000 M-I cm-I at 450 nm, substantially larger than those reported for other iron-nitrosyl complexes. However, if we assume that the majority species is a dinuclear {FeN0}2 complex for which each iron site produces an {FeNO}’-type optical spectrum, we would obtain a reasonable E = 760 M-’ cm-’ per iron, in agreement with values from other iron-nitrosyl complexes (Table 1).

Haskin et al.

11094 Biochemistry, Vol. 34, No. 35, 1995 The {FeN0}7 complex observed here to give rise to an S = 3/2 EPR signal cannot result from a diiron center. If one NO would bind to one iron of R2,,d, the resulting S = 3/2 site would strongly interact with the nearby iron(II), either by exchange interactions as observed for deoxyhemerythrinNO (Nocek et al., 1988) or by strong dipolar interactions that would be recognized with EPR or low-field Mossbauer spectroscopy. Our data are incompatible with either case. Moreover, the high temperature Mossbauer spectra of the sample of Figure 3 did not indicate the presence of any iron(11). This suggests that an {FeN0}7 complex is formed when NO binds to adventitiously bound iron(I1). Since our R2 samples did not contain adventitiously bound iron prior to treatment with NO, some destruction of the diiron center must occur during incubation with NO. The Mossbauer spectra of mononuclear S = 3/2 {FeN0}7 complexes can be analyzed with the S = 3/2 spin Hamiltonian:

x=q+q

I

I

I

1

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

I

y I

0.2

E

I

0.0-

8 o.2 -

w9 0.0-

0.4

-

(1) 0.8 -

with

1. and

where all symbols have their conventional meanings and where go = 2.0. Mossbauer and EPR studies of a variety of {FeN0}7 complexes have revealed a striking similarity of the fine structure and hyperfine structure parameters (D M 10-14 cm-’, E/D 0, Aiso M -34 MHz). The {FeN0}7 minority species observed here has parameters (see caption of Figure 3) very similar to those reported for other NO complexes (Arciero et al., 1983; Bill et al., 1985; Chen et al., 1989; Nocek et al., 1988). We have used this parameter set to simulate the high-field spectra of the {FeN0}7 minority species; the spectra displayed in Figure 4 were obtained by subtracting its simulated contribution (13%) from the raw data. As pointed out above, the Mossbauer spectra of the {FeN0}2 species suggests an electronic system consisting of two exchange coupled S = 3/2 sites. The electronic features of this coupled system can be explored with the Hamiltonian:

+ q-1) + 9 3 2 ) and the x(i) is given by eq 2.

E=JS,*S,

(4)

where SI = S2 = 3/2 Because of the large D values of the mononuclear NO complexes, the electronic properties of the coupled system will be strongly dependent on the zero-field splitting terms. We have investigated the spin expectation values resulting from eq 4, allowing the D tensors of the two sites to be rotated relative to each other. These studies have shown that the magnetic features of the spectra of Figure 4 cannot be explained by assuming ferromagnetic coupling ( J < 0). Thus, for any J < 0, eq 4 would predict, contrary to the experimental observations, the observation of strong magnetic hyperfine interactions in an applied field of 2 T. For antiferromagnetic coupling, two regimes can be distinguished. For J > D, the ground state would essentially be a pure S = 0 level that would yield, at 4.2 K, Mossbauer spectra characteristic of a

I

I

I

I

I

I

I

I

-10 -8 -6 -4

I

I

-2 0 2 4 VELOCnY ( d s )

6

8

10

+

FIGURE 4: 4.2 K MBssbauer spectra of R2,d NO (same sample as that of Figure 3). Spectra were recorded in parallel applied fields as indicated. The spectra were prepared from the raw data by subtracting the theoretical spectra (13% of total absorption) of the monomeric S = 3 / 2 {FeNO}’ complex. The data shown still contain the 10% contribution of the unknown species with AEQ % 1.4 m d s and 6 % 0.07 m d s . The theoretical curves are spectral simulations for a system consisting of two coupled S = 3 / 2 species with J = 5 cm-’, D1 = D2 = 14 cm-’, (E/D)1 = (E/D)2 = 0.02, isotropicA1 = -31.5 MHz and A2 = -35 MHz, AEQ(~) = -1.7 m d s , h E ~ ( 2 ) = -2.1 m d s , rl = q2 = 0, and 61 = 62 = 0.75 m d s . Moreover, the z axes of the two D tensors are tilted by 30° relative to each other. The theoretical curves are plotted to represent the fraction of {FeN0}2, Le., 87% of the data shown. diamagnetic compound. However, the 8.0 T spectrum displays substantial paramagnetic hyperfine structure; therefore, the case J > D can be excluded. This suggests that 0 5 J I 15 cm-’. Keeping all local parameters of the two sites to within 20% of those observed for mononuclear {FeN0}7 species, we have performed spectral simulations allowing J and the relative orientations of the D tensors to vary. The best simulations, shown in Figure 4, were obtained for J = 5 cm-’ for the case where the z axes of both D tensors, i.e., the axes of near axial local symmetry, are tilted 30° relative to each other. All {FeN0}7 complexes studied thus far have E D M 0, suggesting that the NO ligand imposes nearly axial symmetry on the S = 3/2system. It is thus not surprising to find that the individual D tensors of the { FeN0}2 complex should have different principal axis systems. Reactivity of the {FeNO}Z Complex. Changes were observed in both the optical and Mossbauer spectra of R2,,d NO samples upon standing. However, no significant change was observed in its EPR spectrum, indicating that the mononuclear { FeNO}’ component is unaffected. Analysis of the optical spectra taken between 40 and 100 min after initial NO exposure of R2,,d to NO shows a slow conversion to the final spectrum (Figure 5). This conversion is evidenced by the appearance of a new band at 370 nm and a decrease in both the 450 and 620 nm bands with an

+

Biochemistry, Vol. 34, No. 35, 1995 11095

Reaction of NO with Ribonucleotide Reductase 0.50

I

I

I

0.14

0.45

,

,

,

I

,

I

,

0.13

the shoulder at the low-energy band of the doublet in the central portion of the spectrum; for comparison, the spectrum of R2,,, is shown in Figure 6B. Finally, the spectrum contains 14% of the unidentified species with AEQ % 1.4 mm/s and d % 0.07 mm/s; the high-energy line of this doublet is marked by the arrow. A second sample gave essentially identical results. Analysis of a sample of R2red NO frozen 300 min after exposure to NO revealed a composition very similar to that of the 100 min samples; the amount of R2,,, had slightly increased to 47% of the sample (Table 2). The experiments just described show that R2,,, is formed at the expense of the {FeN0}2 complex. To further characterize this conversion, kinetic data were obtained by monitoring the optical absorbance at 620 nm from 20 to 230 min. These data, analyzed by the Kezdy-Swinbourne (Figure 5, inset) and Guggenheim methods (Espenson, 1981), indicated a first-order decomposition process and showed linear fits for five half-lives with a first-order rate constant of 0.013 f 0.001 min-'. As expected, changing the protein concentration did not affect the rate constant derived from the kinetic analyses. The conversion of (FeN0}2 to R2,,, raises a question as to the fate of the bound NO. One mechanistic possibility is that the oxo bridge in the formed R2,,, derived from bound NO accompanied by N 2 0 release. To detect N20, the headspace gas was analyzed by gas chromatography at maximum NO complex concentration (-6 min) and at approximately 100 min after NO addition. The net N20 detected in all samples was due to a combination of that produced by the conversion reaction and by background N20. Control samples showed that background N20 was present in all samples due to a combination of contaminating N20 gas in the NO added to the sample and N20 gas produced by the reaction of NO with trace amounts of dithionite not consumed by the reduction of R2,,,. To correct for this, the background net N20 production (0.47 pmol) was subtracted from the net N20 production (0.79 pmol). The corrected amount of N20 detected in three independent reactions was 0.35, 0.33, and 0.30 equiv per diiron center. Because of the high background level of N20, the error in quantitation is estimated to be 50.06. These values thus correspond to the amount of R2,,, formed (0.4 equiv) in the decomposition of the {FeN0}2 complex detected in the Mossbauer spectrum (Table 2).

+

500

600

550

650

700

800

750

Wavelength (nm)

FIGURE 5: Optical spectra of 2.0 mL of 0.29 mM RZred(50 mM HEPES, pH 7.6, 5% glycerol) plus NO at 40, 50, and 100 min after initial NO exposure; the arrow indicates the reaction progress. For comparison, the dashed line shows the spectrum of R2,,, at the same concentration. The inset shows a Kezdy-Swinboume plot of kinetic data from 20 to 230 min after exposure of 0.5 mL of 0.40 mM R2,,d (50 mM HEPES, pH 7.6, 5% glycerol) to NO.

b

I

-10

I

I

-8

I

I

-6

I

I

-4

I

I

I

I

I

I

I

I

-2 0 2 4 VELOClTY ( m d s )

I

I

6

I

I

8

I

I

I

10

FIGURE6: (A) 4.2 K Mossbauer spectrum of 0.48 mM 57Feenriched R2,,d (50 mM HEPES, pH 7.6,5% glycerol) NO frozen 100 min after exposure to NO. The solid line drawn through the spectrum represents the contribution of the { FeN0}7 complex scaled to 13% of the total absorption. The arrow indicates the high energy line of the quadrupole doublet of the unidentified species. (B) 4.2 K Mossbauer spectrum of RZ,,,.

+

isosbestic point at 440 nm, due to the decay of the {FeN0}2 complex. The spectral evolution suggests that {FeN0}2 converts to a new species that resembles R2,,,. The changes observed in the optical spectra are also evident in the Mossbauer spectra. Figure 6A shows a 4.2 K Mossbauer spectrum of a R2,,d NO sample frozen 100 min after exposure to NO. The spectrum shows that 13% of the iron in the sample belongs to the S = 3/2 {FeNO}' complex, which remains unchanged from the original sample; its contribution is outlined by the theoretical curve drawn through the data. The spectrum exhibits additional contributions from the {FeN0}2 complex (30%) and from R2,,, (42%) (Table 2 ) . The presence of the latter is indicated by

+

DISCUSSION Nitric oxide reacts with the diiron centers of R2,,d to form a mixture of three species. The majority species is an {FeN0}2 complex (77%) which results from the binding of one NO molecule to each iron(I1) of R2redto form a weakly coupled antiferromagnetic pair of {FeNO}' centers (J % 5 cm-'; F=J S I - S ~ ) .The second species is an {FeN0}7 complex (13%) formed when NO binds to adventitious iron(11), while the third species (10%) is of unknown origin. These assignments are primarily based on Mossbauer spectroscopy and supported by optical and EPR evidence; the two nitrosyl complexes observed have spectral properties similar to those of other non-heme mononuclear and dinuclear iron-nitrosyl complexes (Table 1). The weak antiferromagnetic interaction between the {FeN0}7 centers of the {FeN0}2 complex suggests that the

Haskin et al.

11096 Biochemistry, Vol. 34, No. 35, 1995

Po i"

+

FeF -/

'W'

L

_I

170

J

sym(FeNOJ2Complex

\

\

\

r

':J

/

fc

J

["TGEmF]-

1

FIGURE 7: Two possible ways for NO to coordinate to R2,,d to form the {FeN0}2 complex. N-N bond formation would occur only with the syn-bound form as shown. coupling interaction is unlikely to be mediated by hydroxo or oxo bridges. The absence of information on dinuclear [FeN0}7 complexes forces us to make comparisons using the larger database of diiron(II1) and diiron(I1) complexes (Kurtz, 1990; Que & True, 1990). Oxo bridges are only observed in diiron(II1) complexes and lead to coupling constants above 200 cm-', while hydroxo-bridged diiron(111) and diiron(I1) complexes have coupling constants of 2535 cm-' (Kurtz, 1990; Que & True, 1990). On the other hand, @-aqua)diiron(II) complexes with two supporting carboxylate bridges have been synthesized and found to exhibit weak antiferromagnetic coupling [see footnote 7 in Hagen and Lachicotte (1992)l. The presence of this core structure may account for the observed coupling in the [FeN0}2 complex. Weak antiferromagnetic coupling ( 2 cm-I) has also been observed for [Fe,@-02CCH3)2(TPA)21(BPh& (MCnage et al., 1992), a complex which shares the same bis@-carboxylato)diiron(II) core structure as R2red @berg, 1993). Thus, the [FeN0}2 complex could arise from the binding of NO to each of the coordinatively unsaturated iron(I1) centers of R Z d with the core remaining intact (Figure 7). This is also consistent with the observation that two azide ions can bind to the diiron(I1) center at high azide concentrations (Elgren et al., 1993). NO has been very useful as an 0 2 analog to probe the interaction of dioxygen with non-heme iron(I1) centers (Arciero et al., 1983; Chen et al., 1989; Orville & Lipscomb, 1993). The fact that NO reacts with R2,d to afford an [FeN0}2 complex can shed light on the nature of the putative peroxide intermediate in the oxidation of R2,,d to R2,,. It is known that each diiron(I1) center of R2,,d reacts with one dioxygen molecule (Elgren et al., 1991). This initial interaction of dioxygen with R2,d has been proposed to result in the transfer of two electrons from the diiron(I1) center to 0 2 forming a diiron(II1) peroxide intermediate (Fontecave et al., 1990; Ling et al., 1994a; Sahlin et al., 1990). However, this intermediate has not been observed in rapid kinetics studies (Bollinger et al., 1994a,b; Ravi et al., 1994), so the coordination mode of the peroxide in this putative intermediate is not known. The possibilities include terminal ylperoxo, p-y2:y2-peroxo, p-1,l-peroxo, and pu-1,2-peroxo coordination. When NO is used as a probe for 0 2 binding

Intermediate

GlJii, [(pOxo)dliron(lll) Center

FIGURE 8: Proposed partial reaction mechanisms for the reactions of NO and 0 2 with R2,,d. to deoxyhemerythrin, a mono-NO adduct is formed due to the availability of only one exogenous ligand binding site (Stenkamp, 1994). This adduct is described as [Fe"Fe"'NO-] modeling the putative [Fe"Fe"'-02-] intermediate en route to oxyhemerythrin (Nocek et al., 1988). The observation of coupled {FeN0}7 centers in the {FeN0}2 complex of R2 demonstrates that NO can access both iron(I1) ions of the diiron(I1) center and suggests that the bis@-carboxy1ato)diiron core remains intact (Figure 8). The spectroscopic data cannot support a diiron structure with one NO bridging the two iron ions because such a moiety would not have a diamagnetic ground state. In the [FeN0}2 complex, each NO molecule accepts an electron from the diiron(I1) center, modeling the proposed two-electron reduction of 0 2 to peroxide in the initial step of the oxidation of R2,,d. We thus favor a p- 1,l- or a p- 1,2-peroxo mode for the proposed 0 2 adduct of R2,,d. The notion of a bridged peroxide intermediate is further supported by the observed conversion of the {FeN0}2 complex to R2,,, and N20. This conversion is indicated by the appearance of a new species in the Mossbauer spectrum at the expense of the {FeN0}2 complex. The new species has an absorption band at 370 nm and Mossbauer parameters that are consistent with its assignment to R2,,,. Analysis of the kinetics of conversion reveals that this decomposition is a first-order reaction with a rate constant of 0.013 min-'. Furthermore, GC analysis of the reaction vial headspace gas shows that N20 is produced during this conversion in amounts nearly equimolar with the amount of R2,,, formed. The formation of N20 requires N-N bond formation from presumably proximally bound NO molecules. For the [FeN0}2 complex, this would be best accomplished with a syn-bound species as shown in Figure 7. The same available coordination sites that allow a syn-bound {FeN0}2 complex to form would also permit a bridged peroxide intermediate to form. The observation that a significant portion (22%) of the {FeN0}2 complex remains after 300 min suggests that there is a minority population of [FeN0}2 that cannot convert to R2,,, and N20. This minor subset may consist of molecules with the two NO'S bound anti with respect to each other (Figure 7), so that N-N bond formation cannot occur.

Biochemistry, Vol. 34, No. 35, 1995 11097

Reaction of NO with Ribonucleotide Reductase This study is the first example of N20 formation from the reaction of NO with iron(I1) in non-heme iron proteins. Recent spectroscopic studies conclude that the binding of NO to an Fe(I1) center results in the transfer of electron density from the Fe(I1) to the NO forming an Fe(II1)-NOspecies (Brown et al., 1995). The two NO- moieties of the (FeN0}2 complex can then couple to afford an equivalent of N20 and an oxide ion which resides as a bridge between the iron(1II) ions of R2,,,. This reaction can be related to the reductive coupling of two NO molecules that is known to occur in solution (2H' 2e2N0 N20 H20) (Bottomley, 1989). Deoxyhemerythrin-NO does not undergo similar decomposition to produce N20 presumably because only one molecule of NO is present in the active site (Nocek et al., 1988), but it is possible that other dinuclear non-heme iron proteins that can bind 2 equiv of NO can also carry out the conversion of NO to N20. It should be noted that N20 formation has precedence with other proteins and model complexes. For example, the reduced forms of the dicopper protein hemocyanin (Verplaaetse et al., 1979) and the copper-heme protein cytochrome c oxidase (Brudvig et al., 1980) each react with two NO molecules forming dinitrosyl intermediates which decompose to yield N20 and the oxidized proteins. N20 formation has also been demonstrated in the reactions of copper model complexes with NO (Paul & Karlin, 1991). Each copper center reacts with an equivalent of NO to form an unstable intermediate, which decomposes to form a @oxo)dicopper(II) complex accompanied by the release of N20. These results, and the observed decomposition of the R2 {FeN0}2 species, suggest that the formation of the thermodynamically stable oxobridged dimetal center may promote N20 formation. That this is also the driving force in the reaction of R2,d with 0 2 strengthens the argument that NO serves as a good probe of the dioxygen binding site of non-heme iron proteins. The slow conversion of {FeN0}2 to N20 may be due to ratelimiting N-N bond formation. The diiron center understandably may not be as efficient for carrying out such a coupling reaction, since this step is not encountered in the corresponding 0 2 reaction. A mechanism consistent with the results obtained from the majority of the sample in the reaction of NO with R2red is shown in Figure 8 with the proposed reaction of 0 2 with R2,,d. In this mechanism, the {FeN0}2 complex is an intermediate that parallels the proposed bridging diiron(II1) peroxide intermediate of the radical generating reaction. The bridging peroxide intermediate contrasts the q'-hydroperoxo adduct that forms in the reaction of 0 2 with deoxyhemerythrin. This difference in peroxide coordination may play a role in the subsequent reactivity of these intermediates and therefore in overall protein function. To what extent the peroxide coordination mode influences its subsequent reactivity is not presently known and must await information from the investigations of peroxide intermediates of other diiron proteins and model complexes. Both the {FeN0)2 and peroxide intermediates of R2 spontaneously convert to the @-oxo)diiron(III) center; the NO reaction releases N20 in this process while the 0 2 reaction oxidizes T y r m Thus, the reaction of R2red with NO can be viewed as a two-electron analog for the four-electron dioxygen reaction. Although the (p-oxo)diiron(III) center is formed in both reactions, only the dioxygen reaction has the additional oxidizing equivalents to generate the essential Tyr122*.

+

+

-

+

The observed reactivity of R2,,d with NO may contribute to the in vivo inhibition of ribonucleotide reductase by NO. The body's immune system has macrophages that deliver large amounts of NO, via NO synthase, to nearby tumor cells where it can affect several metalloenzymes, one of which is ribonucleotide reductase. It is known that stimulated macrophages produce enough NO to inhibit the enzyme so that the conversion of ribonucleotides into deoxyribonucleotides stops (Feldman et al., 1993). Previously, it has been shown that NO destroys the Tyr122. in tumor cells and in purified bacterial or mammalian R2 proteins, thereby inhibiting ribonucleotide reductase activity (Lepoivre et al., 1991, 1992, 1994). It has been proposed that this inhibition results from Tyr122' quenching by NO. However, the reaction of NO with the diiron(I1) center may provide another mechanism for the inhibition of ribonucleotide reductase. The regeneration of active R2 requires reduction to its R2,,d form and subsequent reaction with 0 2 . NO can oxidize R2,d to form R2,,, without forming the catalytically essential Tyr122.; thus the resulting R1R2 holoenzyme complex will not be capable of reducing the substrate ribonucleotides.

ACKNOWLEDGMENT We thank Professor J. Stubbe for providing E. coli strain N640YpSPS2, Professor John D. Lipscomb for the use of the EPR spectrometer, and Professor William B. Tolman for the use of the gas chromatograph and N20 gas. We thank Professor Michael P. Hendrich for valuable discussions about EPR spectroscopy. We also thank Christy E. Ruggiero and Susan M. Carrier for assistance in N2O quantitation and kinetic data interpretation.

REFERENCES Aberg, A. (1993) Ph.D. Thesis, Stockholm University. Arciero, D. M., Lipscomb, J. D., Huynh, B. H., Kent, T. A., & Munck, E. (1983) J. Biol. Chem. 258, 14981-14991. Barlow, T., Eliasson, R., Platz, A., Reichard, P., & Sjoberg, B.-M. (1983) Proc. Natl. Acad. Sci. U.S.A. 80, 1492-1495. Bill, E., Bernhardt, F.-H., Trautwein, A. X., & Winkler, H. (1985) Eur. J. Biochem. 147, 177-182. Bollinger, J. M., Jr., Tong, W. H., Ravi, N., Huynh, B. H., Edmondson, D. E., & Stubbe, J. (1994a) J. Am. Chem. Soc. 116, 80 15-8023. Bollinger, J. M., Jr., Tong, W. H., Ravi, N., Huynh, B. H., Edmondson, D. E., & Stubbe, J. (1994b) J. Am. Chem. Soc. 116, 8024-8032. Bottomley, F. (1989) Reactions of Coordinated Ligands, Plenum Press, New York. Brown, C. A., Pavlosky, M. A., Westre, T. E., Zhang, Y., Hedman, B., Hodgson, K. O., & Solomon, E. I. (1995) J. Am. Chem. Soc. 117, 715-732. Brudvig, G. W., Stevens, T. H., & Chan, S. I. (1980) Biochemistry 19, 5275-5285. Chen, V. J., Orville, A. M., Harpel, M. R., Frolik, C. A., Surerus, K. K., Munck, E., & Lipscomb, J. D. (1989) J. Bid. Chem. 264, 21677-2168 1. Dawson, J. H. (1988) Science 240, 433-439. Dong, Y., MCnage, S., Brennan, B. A., Elgren, T. E., Jang, H. G., Pearce, L. L., & Que, L., Jr. (1993) J. Am. Chem. SOC.115, I85 1- 1859. Elgren, T. E., Lynch, J. B., Juarez-Garcia, C., Munck, E., Sjoberg, B.-M., & Que, L., Jr. (1991) 1.Biol. Chem. 266, 19265-19268. Elgren, T. E., Hendrich, M. P., & Que, L., Jr. (1993) J. Am. Chem. Soc. 115, 9291-9292. Enemark, J. H., & Feltham, R. D. (1974) Coord. Chem. Rev. 13, 339-406. Espenson, J. H. (1981) Chemical Kinetics and Reaction Mechanisms, McGraw-Hill, Inc., New York.

11098 Biochemistry, Vol. 34, No. 35, 1995 Feig, A. L., & Lippard, S . J. (1994) Chem. Rev. 94, 759-805. Feldman, P. L., Griffity, 0. W., & Stuehr, D. J. (1993) Chem. Eng. News 71, 26-38. Fontecave, M., Gerez, C., Atta, M., & Jeunet, A. (1990) Biochem. Biophys. Res. Commun. 168, 659-664. Guengerich, F. P. (1991) J. Biol. Chem. 266, 10019-10022. Hagen, K. S., & Lachicotte, R. (1992) J. Am. Chem. SOC. 114, 8741-8742. Kennedy, M. C., Gan, T., Antholine, W. E., & Petering, D. H. (1993) Biochem. Biophys. Res. Commun. 196, 632-635. Kitajima, N., & Moro-oka, Y . (1994) Chem. Rev. 94, 737-757. Kitajima, N., Tamura, N., Amagai, H., Fukui, H., Moro-oka, Y . , Mizutani, Y . , Kitagawa, T., Mathur, R., Heerwegh, K., Reed, C. A., Randall, C. R., Que, L., Jr., & Tatsumi, K. (1994) J. Am. Chem. SOC.116, 9071-9085. Kurtz, D. M., Jr. (1990) Chem. Rev. 90, 585-606. Larsson, A., & Sjoberg, B.-M. (1986) EMBO J. 5, 2037-2040. Lee, S.-K., Nesheim, J. C., & Lipscomb, J. D. (1993) J. Biol. Chem. 268, 21569-21577. Lepoivre, M., Fieschi, F., Coves, J., Thelander, L., & Fontecave, M. (1991) Biochem. Biophys. Res. Commun. 179, 442-448. Lepoivre, M., Flaman, J.-M., & Henry, Y. (1992) J. Biol. Chem. 267, 22994-23000. Lepoivre, M., Flaman, J.-M., BobC, P., Lemaire, G., & Henry, Y. (1994) J. Biol. Chem. 269, 21891-21897. Ling, J., Nestor, L. P., Czernuszewicz, R. S . , Spiro, T. G., Fraczkiewicz, R., Sharma, K. D., Loehr, T. M., & Sanders-Loehr, J. (1994a) J. Am. Chem. SOC.116, 7682-7691. Ling, J., Sahlin, M., Sjoberg, B.-M., Loehr, T. M., & SandersLoehr, J. (1994b) J. Biol. Chem. 269, 5595-5601. Lynch, J. B. (1989) Ph.D. Thesis, University of Minnesota, Minneapolis, MN. Lynch, J. B., Juarez-Garcia, C., Miinck, E., & Que, L., Jr. (1989) J. Biol. Chem. 264, 8091-8096. Magnus, K. A., Ton-That, H., & Carpenter, J. E. (1994) Chem. ' R e v . 94, 727-735. MCnage, S., Zang, Y., Hendrich, M. P., & Que, L., Jr. (1992) J. Am. Chem. SOC.114, 7786-7792.

Haskin et al. Nelson, M. J. (1987) J. Biol. Chem. 262, 12137-12142. Nocek, J. M., Kurtz, D. M., Jr., Sage, J. T., Xia, Y.-M., Debrunner, P., Shiemke, A. K., Sanders-Loehr, J., & Loehr, T. M. (1988) Biochemistry 27, 1014-1024, Nordlund, P., & Eklund, H. (1993) J. Mol. Biol. 232, 123-164. Nordlund, P., Sjoberg, B.-M., & Eklund, H. (1990) Nature 345, 393-398. Orville, A. M., & Lipscomb, J. D. (1993) J. Biol. Chem. 268,85968607. Paul, P. P., & Karlin, K. D. (1991) J. Am. Chem. SOC.113,63316332. Paul, P. P., Tyeklh, Z., Jacobson, R. R., & Karlin, K. D. (1991) J. Am. Chem. SOC. 113, 5322-5332. Que, L., Jr. (1991) Science 253, 273-274. Que, L., Jr., & True, A. E. (1990) Prog. Inorg. Chem. 38, 97200. Ravi, N., Bollinger, J. M., Jr., Huynh, B. H., Edmondson, D. E., & Stubbe, J. (1994) J. Am. Chem. SOC.116, 8007-8014. Reichard, P. (1993) Science 260, 1773-1777. Sahlin, M., Sjoberg, B.-M., Backes, G., Loehr, T. M., & SandersLoehr, J. (1990) Biochem. Biophys. Res. Commun. 167, 813818. Salowe, S . P. (1987) Ph.D. Thesis, University of Wisconsin, Madison, WI. Salowe, S . P., & Stubbe, J. (1986) J. Bacteriol. 165, 363-366. Shiemke, A. K., Loehr, T. M., & Sanders-Loehr, J. (1986) J. Am. Chem. SOC. 108, 2437-2443. Stenkamp, R. E. (1994) Chem. Rev. 94, 715-726. Stubbe, J. (1990) Adv. Enzymol. Relat. Areas Mol. Biol. 63, 349419. Uhlin, U., & Eklund, H. (1994) Nature 370, 533-539. Verplaaetse, J., van Tornout, P., Defreyn, G., Witters, R., & Lontie, R. (1979) Eur. J. Biochem. 95, 327-331. BI9.5073 1P